next up previous
Next: Classical Statistical Mechanics Up: Statistical Mechanics: A Brief Previous: Making Observations: The Ergodic

Entropy and Temperature

Eq. 2.1 introduced the quantity $\Omega(N,V,E)$ as the number of states available to a system under the constraints of constant number of particles, $N$, volume, $V$, and energy $E$. The fundamental postulate of statistical mechanics, also called the ``rational basis'' by Chandler, is the following:

In statistical equilibrium, all states consistent with the constraints of $N$, $V$, and $E$ are equally probable.
or
\begin{displaymath}
P_\nu = 1/\Omega(N,V,E).
\end{displaymath} (9)

This relation is often referred to as a statement of the ``equal a priori probabilities in state space.'' Another way of saying the same thing: The probability distribution for states in the microcanonical ensemble is uniform.

The link between statistical mechanics and classical thermodynamics is given by a definition of entropy:

\begin{displaymath}
S \equiv k_B\ln\Omega
\end{displaymath} (10)

Note two important properties of $S$. First, it is extensive: if we consider a compound system made of subsystems $A$ and $B$ with $\Omega_A$ and $\Omega_B$ as the respective number of states, the total number of states is $\Omega_A\Omega_B$, and therefore $S = S_A +
S_B$. Second, it is consistent with the second law of thermodynamics: putting any constraint on the system lowers its entropy because the constraint lowers the number of accessible states.

Temperature is defined using entropy: $1/T \equiv \left(\partial S/\partial E\right)_{N,V}$, or

\begin{displaymath}
\beta = \left(k_BT\right)^{-1} = \left(\partial\ln\Omega/\partial E\right)
\end{displaymath} (11)

Now we will consider constraining our system not with constant $E$, but with constant $T$. The set of all possible states satisfying constraints of $N$, $V$, and $T$ is called the canonical ensemble. Contrary to appearances based on their names, the canonical ensemble can be envisioned as a subsystem in a larger, microcanonical system. Consider such a system divided into a small subsystem $A$, surrounded by a large ``bath'' $B$. We imagine that these two subsystems are ``weakly coupled,'' meaning they exchange only thermal energy, but no particles, and their volumes remain fixed. We seek to compute the probability of finding the total system in a state such that subsystem $A$ has energy $E_A$. The entire system is microcanonical, so the total energy, $E$ is constant, as is the total number of states available to the system, $\Omega (E)$ (we omit the $N$ and $V$ for simplicity).

When $A$ has energy $E_A$, the total system energy is $E = E_A + E_B$, where $E_B$ is the energy of the bath. By constraining system $A$'s energy, we have reduced the number of states available to the whole system to $\Omega(E-E_A)$. So, the probability of observing the system in a state in which subsystem $A$ has energy $E_A$ looks like

\begin{displaymath}
P_A = \frac{\left[\begin{array}{l}
\mbox{Number of states f...
...ht]} = \frac{\Omega\left(E - E_A\right)}{\Omega\left(E\right)}
\end{displaymath} (12)

We can expand $\Omega\left(E - E_A\right)$ in a Taylor series around $E_A = 0$:
$\displaystyle P_A \propto \Omega\left(E - E_A\right)$ $\textstyle =$ $\displaystyle \exp\left[\ln\Omega\left(E-E_A\right)\right]$ (13)
  $\textstyle =$ $\displaystyle \exp\left[\ln\Omega\left(E\right) - E_A\frac{\partial\ln\Omega}{\partial E} + \cdots\right],$ (14)

where the partial derivative implies we are holding $N$ and $V$ fixed. We can truncate the Taylor expansion at the first-order term, because higher order terms become less and less important as the size of subsystem $B$ becomes larger and larger. What results is the Boltzmann distribution law for energies of a system at constant temperature:
\begin{displaymath}
P_A \propto \exp\left(-\beta E_A\right)
\end{displaymath} (15)

The normalization condition requires that for all energies of subsystem $A$, $E_A$,
\begin{displaymath}
\sum_A P_A = 1 = \sum_A \exp\left(-\beta E_A\right) \equiv Q\left(N,V,T\right),
\end{displaymath} (16)

which defines the canonical partition function, $Q$. Therefore,
\begin{displaymath}
P_A = Q^{-1}\exp\left(-\beta E_A\right)
\end{displaymath} (17)

Because some energies can correspond to more than one microstate, we should distinguish between ``states'' and ``energy levels.'' We can express the canonical partition function as

\begin{displaymath}
Q = \sum_{\begin{array}{cc}\nu\ \mbox{states}\end{array}} e...
...mbox{levels}\end{array}} \Omega\left(E_l\right) e^{-\beta E_l}
\end{displaymath} (18)

where, as we have seen, $\Omega\left(E\right)$ is the number of microstates with energy $E$. Moving to the continuum limit, and assuming a reference energy of $E_{ref} = 0$,
\begin{displaymath}
Q \rightarrow \int_0^\infty dE \overline\Omega\left(E\right)e^{-\beta E}
\end{displaymath} (19)

where $\overline\Omega\left(E\right)$ is the density of states. What is this equation telling us? It is telling us that $Q$ is the Laplace transform of $\overline\Omega$. We know that transform pairs are unique, and hence, both $Q$ and $\overline\Omega$ contain the same information.

We recognize that for a system described by a canonical ensemble, the energy is a fluctuating quantity. And we now have the probability of observing a state with a given energy, so we can use Eq. 1 to compute the average energy, $\left<E\right>$. Consider

$\displaystyle \left<E\right>$ $\textstyle =$ $\displaystyle \left<E_\nu\right> = \sum_\nu P_\nu E_\nu$ (20)
  $\textstyle =$ $\displaystyle \left[\sum_\nu E_\nu \exp\left(-\beta E_\nu\right)\right] \left/ \left[\sum_\nu \exp\left(-\beta E_\nu\right)\right] \right.$ (21)

Notice that
\begin{displaymath}
\sum_\nu E_\nu \exp\left(-\beta E_\nu\right) = -\left(\partial Q \left/ \partial\beta\right.\right)
\end{displaymath} (22)

Recalling that $d\ln f(x) / dx = 1/f df/dx$, we see that
$\displaystyle \left<E\right>$ $\textstyle =$ $\displaystyle -\left(\partial Q \left/ \partial\beta\right.\right)\left/Q\right.$ (23)
  $\textstyle =$ $\displaystyle -\left(\partial \ln Q \left/ \partial\beta\right.\right)_{N,V}$ (24)

Now, let us consider the average magnitude of the fluctuations in energy in the canonical ensemble.

$\displaystyle \left<\left(\delta E\right)^2\right>$ $\textstyle =$ $\displaystyle \left<\left(E - \left<E\right>\right)^2\right>$ (25)
  $\textstyle =$ $\displaystyle \left<E^2\right> - \left<E\right>^2$ (26)
  $\textstyle =$ $\displaystyle \sum_\nu P_\nu E_\nu^2 - \left(\sum_\nu P_\nu E_\nu\right)^2$ (27)
  $\textstyle =$ $\displaystyle Q^{-1}\left(\frac{\partial^2Q}{\partial\beta^2}\right)_{N,V} -
Q^{-2}\left(\frac{\partial Q}{\partial\beta}\right)^2_{N,V}$ (28)
  $\textstyle =$ $\displaystyle \left(\frac{\partial\ln Q}{\partial\beta^2}\right)_{N,V} = -\left(\frac{\partial\left<E\right>}{\partial\beta}\right)_{N,V}$ (29)

Now, noting that the definition of heat capacity at constant volume, $C_v$, is
\begin{displaymath}
C_v = \left(\frac{\partial E}{\partial T}\right)
\end{displaymath} (30)

we see that
\begin{displaymath}
\left<\left(\delta E\right)^2\right> = k_B T^2 C_v
\end{displaymath} (31)

This is an interesting statement. It relates the magnitude of spontaneous fluctuations in the total energy of a system to that system's capacity to store or release energy due to changing its temperature.

The fact (Eq. 24) that the average energy in the canonical ensemble is related to a derivative of the log of the partition function implies that $\ln Q$ is an important thermodynamic quantity. So, let's go back to our undergraduate thermodynamics course(s) and recall the following statement of the 1st and 2nd Law:

\begin{displaymath}
dA = -SdT - pdV + \mu dN
\end{displaymath} (32)

where $A$ is the Helmholtz free energy, defined in terms of internal energy and entropy as
\begin{displaymath}
A = \left<E\right> - TS
\end{displaymath} (33)

Now, consider the following derivative of $A$:
$\displaystyle \left(\frac{\partial\left(A/T\right)}{\partial\left(1/T\right)}\right)_{N,V}$ $\textstyle =$ $\displaystyle A + \frac{1}{T}\left(\frac{\partial A}{\partial\left(1/T\right)}\right)_{N,V}$ (34)
  $\textstyle =$ $\displaystyle A - T\left(\frac{\partial A}{\partial T}\right)_{N,V} = A - TS = \left<E\right>.$ (35)

Therefore,
\begin{displaymath}
\left(\frac{\partial\left(\beta A\right)}{\partial\beta}\right)_{N,V} = \left<E\right>.
\end{displaymath} (36)

Considering Eq. 24, we see that
\begin{displaymath}
\ln Q + C = -\beta A
\end{displaymath} (37)

which does indeed suggest an important link between $\ln Q$ and the important thermodynamic quantity, the Helmholtz free energy. But what is the constant $C$? To evaluate it, consider the ``boundary condition'' as $T \rightarrow 0$:
\begin{displaymath}
Q = \sum_\nu e^{-\beta E_\nu} \stackrel{T\rightarrow 0}{\longrightarrow} e^{-\beta E_{\rm ground}}.
\end{displaymath} (38)

Here, we have assumed that the degeneracy of the ground state, $\Omega\left(E_{\rm ground}\right)$ is 1. This tells us that
\begin{displaymath}
\lim_{T\rightarrow 0} \ln Q = -\beta E_{\rm ground}
\end{displaymath} (39)

Using this fact, and combining Eqs. 33 and 37, as $T \rightarrow 0$, we see that
\begin{displaymath}
-\beta E_{\rm ground} + C = -\beta \underbrace{\left<E\right...
...}} - \underbrace{\frac{S}{k_B}}_{\rightarrow 0 (\Omega = 1)},
\end{displaymath} (40)

Hence, $C$ = 0. So,
\begin{displaymath}
\ln Q = -\beta A.
\end{displaymath} (41)

The quantity $-\beta^{-1}\ln Q$ is the Helmholtz free energy, $A$. This is denoted $F$ in Frenkel & Smit [1].


next up previous
Next: Classical Statistical Mechanics Up: Statistical Mechanics: A Brief Previous: Making Observations: The Ergodic
cfa22@drexel.edu